View/Open - University College Cork

advertisement
Application of Electrochemical Impedance for Characterising
Arrays of Bi2S3 Nanowires
Gunta Kunakova,1 Juris Katkevics,1,2 Arturs Viksna,1,2 Zanda Gertnere,1 Justin Varghese,3,4
Justin D. Holmes3,4 and Donats Erts1
1
Institute of Chemical Physics, University of Latvia, Raina Blvd. 19, Riga, LV-1586, Latvia
Department of Analytical Chemistry, University of Latvia, Kr. Valdemara 48, LV-1013, Latvia
3
Materials Chemistry & Analysis Group, Department of Chemistry and the Tyndall National Institute,
University College Cork, Ireland
4
Centre for Research on Adaptive Nanostructures and Nanodevices (CRANN), Trinity College Dublin,
Dublin 2, Ireland
2
E-mail: donats.erts@lu.lv
Abstract. Electrochemical Impedance Spectroscopy (EIS) was used to characterise the electrical properties
of bismuth sulphide (Bi2S3) nanowires (NWs) templated within anodic aluminium oxide (AAO) membranes.
A specially engineered cell, with a nominal electrolyte volume of 0.1 – 0.2 ml, was used to hold and measure
the electrochemical impedance of the fragile NW/AAO samples. An equivalent circuit model was developed
to determine the filling density of nanowires within the porous templates. The EIS method can be utilised to
probe the nanowire filling density in porous membranes over large sample areas, which is often unobtainable
using electron microscopy and conductive atomic force microscopy techniques.
1. Introduction
Bi2S3 is a semiconductor with a direct band gap of 1.3 eV which has been applied as an active
material in biomolecule detection [1], gas sensing [2,3] and optoelectronic devices [4]. These
applications have also included the incorporation of Bi2S3 nanowires into ordered arrays using
porous templates Anodic aluminium oxide (AAO) is one of the most frequently used templates
employed to produce arrays of nanowires in combination with electrochemical, supercritical fluid
and chemical vapour deposition approaches [5-8]. Several characterisation methods have been
employed to determine the electrical properties and nanowire filling densities within porous
membranes [5,7,9-13]. In particular, gas adsorption methods have been utilised to characterise the
volume of empty nanopores within templating membranes [9] and electron microscopy techniques
have been developed to characterise the structure and filling density of nanowire arrays at a local
scale [5,7,8]. Conductive atomic force microscopy (C-AFM) has also been applied to determine the
quality and density of the nanowires inside porous templates [10-13], although information about
partly filled nanopores cannot be extracted from these data. C-AFM is also inadequate for the large
scale characterisation of samples.
Electrochemical impedance spectroscopy (EIS) is widely used for characterising the
properties of large area electrodes, including those coated with porous layers [14]. EIS methods
have been used previously to characterise the electrical behaviour of free standing nanowires for
different applications, such as electrochemical capacitors [15], photo-anodes [16] and elements for
photovoltaics [17]. EIS has also been applied for resolving the electrodeposition parameters of
metallic nanowires into the pores of AAO membranes and their interfacial electron transport [1820]. However, to-date there has been no reports about the application of EIS for characterising the
electrical properties and filling densities of nanowires templated within porous membranes, such as
AAO, over the large areas. In this article, we detail an EIS-based method for the comprehensive and
non-destructive characterisation of templated complex nanomaterials, such as arrays of assembled
nanowires. Here we detail the characterisation of arrays of Bi2S3/AAO nanowire assemblies,
through the construction of equivalent circuits and electrochemical cells, although our approach can
also be applied to the analysis of the other types of templated nanomaterials.
1
2. Experimental
2. 1. Sample Preparation
Arrays of Bi2S3 nanowires were fabricated inside the pores of AAO membranes, with nominal
diameters of 200 nm, by the thermolysis of the single source precursor bismuth
bis(diethyldithiocarbamate) [Bi(S2CNEt2)3]. The synthesis method for the Bi2S3 nanowires into the
AAO membranes detailed is described in the Ref. [21]. Structural characterisation of the nanowire
arrays were performed using X-ray diffraction (XRD), high resolution transmission electron
microscopy (TEM), selected area electron diffraction SAED as described in detail elsewhere [21].
Initially both surfaces of the AAO membranes containing the nanowires were polished to remove
reaction products from the surface. Polishing of the membranes was performed using the diamond
suspensions (Buehler) with different size of the particles for each suspension. For the initial
polishing a suspensions with the particle size of 6; 1; 0.5 µm were used and for the final step: 0.1
and 0.05 µm respectively. Au was then deposited on one side of the polished membranes to provide
electrical contact to the nanowires within the templates. Arrays of the nanowires were visualised by
scanning electron microscopy (SEM) (Hitachi S4800) at an acceleration voltage of 2 kV. The
number of empty/filled pores per cm2 was determined from the SEM images.
Conductive atomic force microscopy (C-AFM) measurements were performed using an
Asylum Research MFP-3D instrument. Commercially available diamond coated DCP20 tips from
NT-MDT with a nominal radius of 100 nm were used in all measurements. The C-AFM tips were
grounded whilst a 10 V dc bias was applied to the sample. The topographic images of the samples
were recorded in contact mode simultaneously with the current image of the same area. Energy
dispersive X-ray maps were obtained using Bruker Quantax spectrometer installed on a scanning
electron microscope (SEM) (Hitachi S4800) operating at accelerating voltages of 25 and 30 kV.
2. 2. Electrical and Electrochemical Impedance Measurements
Samples were integrated in electrolytic cells with a total volume of between 0.1 – 0.2 ml
(Figure 1). For electrochemical measurements, Bi2S3/AAO samples were used as cathodes, Pt
plates were used as anodes and Ag/AgCl was used as the reference microelectrode. A 0.1 M
solution of sodium acetate (CH3COONa) was used as the cell electrolyte. Obtaining reliable and
accurate EIS data required precise adjustment (0.25 – 0.35 cm) of the distance between the cathode
and anode. Adjustment was performed using a micrometric screw under a binocular microscope.
Before any electrochemical impedance measurements were undertaken, each sample was
conditioned (immersion) in the electrolyte solution for 30 min. All measurements were performed
with an "Autolab PGSTAT 30" potentiostat. The electrochemical characteristic window, kinetics
and potential areas of ion mass transfer for electrochemical impedance measurement were
determined from cyclic voltammograms. Impedance spectra were obtained at potentials of -0.05 V
to -0.30 V and at scan frequencies ranging from 10 kHz to 0.02 Hz, using the 10 mV amplitude of
sinusoidal voltage. Empty AAO membranes were measured in a similar manner. One sample was
used to obtain several EIS scans; but fresh electrolyte was exchanged between the measurements of
each spectrum
2. 3. Analysis of Impedance Spectra
The equivalent electrical circuit was derived and impedance element values were calculated
using frequency response analysis (FRA). Impedance data from the Bi2S3/AAO samples and empty
AAO membranes were analysed using a complex plane Nyquist plot [22-24]. After successful
validation of the experimental data, a Nyquist plot diagram was used to evaluate the number of
elements in the Randle’s equivalent circuit. Parameter fitting for the equivalent circuit elements
was performed by fitting and simulation programs “RCNTRANS” and “EQUIVCRT” definition of
Dr. B. A Boukamp (The FRA Windows), which are based on a non-linear least, squares methods
[25].
3. Results and Discussion
2
Investigation of the electrochemical impedance parameters of arrays of Bi2S3 nanowires
within the AAO membranes was conducted on samples with low (L) and high (H) pore filling
densities. Filling densities, i.e., the percentage of pores within the membranes containing Bi2S3
nanowires were determined from SEM images, over different sample areas as shown in Figure 2. In
the low density nanowire samples, 73-77 % of the membrane pores were filled (Figure 2a),
compared to the high density nanowire samples, where 97-99 % of the nanopores were filled (see
example in Figure 2b).
The number of filled nanopores per sample area was taken into account in the electrochemical
impedance parameter calculations. Results for the samples shown in the paper are summarised in
the Table 1. However, it should be noted that SEM images taken of the samples did not cover the
full sample area, i.e. 1 cm2. SEM analysis cannot be used to evaluate the degree of pore filling, i.e.
pores which are only partially filled with nanowires do not run continuously through the entire
length of the pores.
C-AFM images in the localised areas of the arrays of the nanowires, with a filling density of
approximately 99 %, highlighted the fact that not all of the nanowires seen in the topography
images (Figure 2c) were conductive (Figure 2d) and hence did not run continuously through the full
length of the nanopores. Furthermore energy dispersive X-ray mapping over the high density
nanowire array samples were performed to support the presence of nanowires in the AAO
templates. An example of an EDS map for a high density Bi2S3/AAO nanowire sample is shown in
Figure 2e, confirming that the bright spots in the SEM image refer to the positions of the Bi 2S3
nanowires.
EIS measurements were conducted on nanowire samples and AAO membranes with the aim
of obtaining information about the electrical properties of the nanowires over large areas. In order
to establish the impedance measurement region, the electrochemical window was determined by
measuring the cyclic voltammetric curves for arrays of Bi2S3/AAO nanowire samples and empty
AAO membranes (Figure 3a). The equilibrium potential for both high and low density filled
Bi2S3/AAO samples was found to be between -0.30 to -0.05 V. The slope of the equilibrium
potential zone decreased from 6.67×10-7 V/dec for empty AAO membranes down to around
1.27×10-7 V/dec for the Bi2S3/AAO samples with 77 % filling density, and remained practically
constant for the samples with filling density of 99 %.
Electrochemical impedance spectra of the Bi2S3/AAO nanowire samples were obtained in the
equilibrium potential range of the electrochemical window. To characterise the oxidation transition
equilibrium point, the cyclic voltammetric data of the Bi2S3/AAO nanowire arrays and AAO
nanoporous samples were analysed by the Tafel slope test [14, 25] (Figure 3b, c). For a selected
region NOVA software was used to calculate the Tafel slopes and the corrosion currents. The
overpotential (η) was defined as the difference between the applied potential and the corrosion
potential (Ecorr). The corrosion potential is the open circuit potential for a corroding Bi 2S3 nanowire
or AAO nanoporous wall. For large cathode overpotential (bc < -1), the Tafel equation [14, 25] for a
cathodic reaction is given by equation 1:
η = log icorr. - bc logi
(1)
Ecorr for all Bi2S3/AAO samples and AAO membranes was -0.12 V and -0.24 V respectively. The
corrosion current density (icorr) was 7.310-7A cm-2 and 9.210-7 A cm-2 for all Bi2S3/AAO and AAO
membranes respectively. The calculated values of coefficient bc were 0.084 V and 0.181 V for the
Bi2S3/AAO nanowire arrays and empty AAO membranes respectively.
Nyquist diagrams for empty and Bi2S3 filled AAO samples are shown in the Figure 4 (curves
1 – 3). For the empty AAO membrane, the impedance spectra were similar and comparable at
voltage-s between -0.10 to -0.20 V, for all of the frequency intervals between 10 kHz to 0.02 Hz.
However, for the Bi2S3/AAO samples the impedance spectra were similar and comparable at high
frequencies (10.0 to 1.5 kHz), but different at low frequencies (0.1 Hz to 0.02 Hz). Differences in
the impedance spectra were more pronounced at high filling densities (Figure 5, curve 3) and high
voltages. The difference in the impedance spectra at low frequencies and at different voltages has
previously been observed for Bi2S3 nanoparticles [26]. Since impedance spectra at low frequencies
3
for the AAO membranes filled with Bi2S3 nanowires exhibited similar behavior to Bi2S3
nanoparticles, this indicates that variances in the EIS spectra at low frequencies for filled AAO
membranes is due to the presence of Bi2S3 inside the nanopores.
At the beginning EIS data for all samples were analysed by widely used mathematical equivalent
circuit diversiform (models, formula) for porous electrodes, i.e. circuit description 1 – R(RQ), 2 –
R(RQ)(RQ) and 3 – R([R(RQ)]Q) [14, 25], composed from a combination of resistances (R) and
capacitances (Q). For the empty AAO membranes, only one kink was present in the Bode plots
(Figure 5, curve 4), indicating the presence of a single RQ component. Hence only an R(RQ)
circuit was applied in further analysis.
Errors of calculated element values (Table 2) of equivalent circuit R(RQ) (Figure 6a) for the
empty AAO membranes were below 1 %, indicating that for the empty AAO membranes simple
R(RQ) circuit can be applied to the EIS spectra analysis.
The presence of two kinks in the Bode plots (Figure 5, curves 5, 6) for Bi 2S3 nanowire arrays with
both high and low filling densities confirmed the existence of two parallel RQ structures and
therefore both R(RQ)(RQ) and R([R(RQ)]Q) [14, 25] circuits were considered for EIS spectra
analysis.
Table 3 summarizes fitting parameters of the equivalent circuit R(RQ)(RQ). Estimated errors for
some of the parameters were 100 % for both high and low density nanowire arrays, confirming that
this equivalent circuit could not be used for EIS data analysis. Although errors of R([R(RQ)]Q)
circuit elements were lower in comparison to an R(RQ)(RQ) circuit, an error of 38 % for the
electrolyte resistance (Re) was too large for EIS data analysis of the Bi2S3nanowire arrays.
We have optimised the equivalent circuit of porous electrodes [14] for our system by ordering
elements of the circuit Re(Cℓ1[Rℓ1(ZfCdℓ)]) to Re([R2(R3Q2)]Q1) (Figure 6, b, c), where Re is the
electrolyte resistance; R2 is the resistance in the AAO nanopores, which is the combination of the
resistance from the electrolyte in the nanopores and the resistance of the nanowires; R3 and Q2 is the
resistance and capacitance of the double layer between the electrolyte/Bi2S3 nanowire samples and
the Au electrode respectively; C1 is the capacitance of the AAO pore walls. As can be seen from
the fitted data presented in Table 4, errors of all circuit Re([R2(R3Q2)]C1) elements were below 8 %,
hence this model can be applied for the analysis of arrays of Bi2S3 nanowires with different filling
densities.
The value of R3, which characterises the resistance of the double layer between the nanowires
in the nanopores and the Au electrode increases from 0.54 MΩ cm2 for a low density Bi2S3/AAO
sample, up to 1.03 MΩ cm2 for a high density sample. This increase in the double layer resistance
is probably caused by the formation of a Schottky barrier between the Bi2S3 and Au surfaces. The
value of R2, which characterises the total resistance of a Bi2S3/AAO nanowire sample, increases as
a function of nanowire filling density, i.e. from 1.12 k cm2 up to 2.01 k cm2, for low and high
density samples, respectively. According to B. A. Boukamp classification of capacitance [27], the
capacitance values between 1 – 10 nF∙cm-2 signifies the presence of grain boundaries within the
nanowires, whilst values between 0.1 – 10 μF∙cm-2 are characteristic of a double layer/space
charges. Hence, a capacitance (C1) value of 10 nF∙cm-2 for the low and 60 nF∙cm-2 for the high
density arrays of the nanowires indicates that the ends of the Bi2S3 nanowires act as grain
boundaries. A calculated value for the element Q2, which is characteristic of the capacitance of the
double layer between the Bi2S3 nanowires and the Au electrode (Table 4), was 41.2 μF∙cm-2 for low
density nanowire arrays, decreasing by several orders of magnitude for high density nanowire
arrays, e.g. 0.97 μF cm-2 for 99 % filled membranes.
Bode plots diagrams (Figure 5) were also analysed to characterise the electrochemical
parameters of the materials. These diagrams for empty AAO membranes and Bi2S3/AAO nanowire
arrays showed that EIS spectra could be measured at an applied potential of -0.20 V and below.
EIS spectra could also be measured at -0.1 V for empty AAO samples. The changes in the
impedance phase angle as a function of frequency for empty AAO membranes was fitted with one
time constant, whereas for the low and high density arrays of nanowires could be fitted with two
time constants (f1,max and f2,max) (Figure 5a). The slope values fRC and R+Re were calculated for the
4
empty AAO membrane and all of the arrays of Bi2S3/AAO nanowires. The slope of the impedance
modulus │Z│ frequency at the transition point, on a log – log scale (Figure 5a), was between -0.72
– -0.74 for empty AAO membranes and for the low and high density Bi2S3 nanowire samples. The
time constant for the empty AAO nanowire arrays fRC = (2pRC)-1 = 0.38 – 0.42 s-1 was identical
with one of the time constants for the low and high density Bi2S3/AAO samples (Figure 5a).
Despite the different structure of the materials, for the arrays of Bi2S3 nanowires large volumes of
empty space were filled with the electrolyte. The second time constant (Figure 5b) determined for
the Bi2S3 nanowire arrays fRC= 0.29
s-1 was characteristic of the Bi2S3 material.
Maximal impedance modulus (Re + R) at low frequencies were related to the reactive circuit
with values of 0.04 M cm2, 0.37 – 0.42 M cm2 and 0.50 – 0.87 M cm2 for the empty AAO,
low and high density nanowire arrays respectively (Figure 5a). The rising impedance modulus with
increasing nanowire filling density can be explained by changes in the sample structure containing a
mixture of four components: electrolyte, Bi2S3 nanowires, AAO pore walls and the nanopores
themselves. The relationship between these components changes the impedance. The increase of
the impedance modulus by increasing nanowire filling density can be used to quantify the filling
density over large areas of nanowire arrays grown in nanoporous templates.
To summarize, simple method for evaluation of the pore filling density of the nanoporous samples
is based on determination and comparing of impedance modulus between unknown sample with
impedance modulus for 2 standarts - empty and fully filled nanoporous sample. In absence of
standards difference in filling levels can be determined by comparison of impedance modulus for
different samples.
4. Conclusions
Experimental data has shown that the EIS parameters of a cell containing an AAO membrane
or an array of AAO/Bi2S3 nanowires can be interpreted in terms of the equivalent circuit denoted as
Re([R2(R3Q2)]C1), where Re is the electrolyte resistance; R2 is the resistance in the AAO nanopores
(obtained from the resistance of the electrolyte in nanopores and resistance of the nanowires); R 3
and Q2 are the resistance and capacitance of the double layer between the electrolyte/Bi 2S3
nanowire samples and the Au electrode respectively; C1 is the capacitance of the AAO pore walls.
The values of equivalent circuit and the impedance modulus have been successfully used to
characterise the filling density of arrays of Bi2S3 nanowire within an AAO membrane.
Electrochemical impedance spectroscopy could be considered as an alternative method for
determining the filling density and resistance of templated nanowires arrays, over large areas, which
is difficult to achieve using other techniques. The approach could also be utilised for characterising
arrays of other nanomaterials, e.g. nanotubes, within other templates, e.g. polymer membranes. In
order to characterize other templated nanomaterials with presented approach, one has to consider
several aspects such as compatibility of materials with the electrolyte and formation of the electrical
contact to the nanomaterial sample. Depending on material properties increase of frequency interval
may be required.
5. Acknowledgements
This work has been supported by ERAF project No. 2010/0251/2DP/2.1.1.1.0/10/APIA/VIAA/096.
GK acknowledges the European Social Fund within the project «Support for Doctoral Studies at
University of Latvia». JDH and JV acknowledge financial support from Science Foundation
Ireland (SFI) (Grant: 09/IN1/I2602).
6. References
5
[1] L. Cademartiri, F. Scotognella, P.G. O’Brien, B.V. Lotsch, J. Thomson, S. Petrov, N.P. Kherani, G.A.
Ozin, Cross-linking Bi2S3 ultrathin nanowires: a platform for nanostructure formation and biomolecule
detection, Nano Lett. 9 (2009) 1482.
[2] H. Li, J. Yang, J. Zhang, M. Zhou, Facile synthesis of hierarchical Bi2S3 nanostructures for photodetector
and gas sensor, RSC Adv. 2 (2012) 6258.
[3] A.D. Schricker, M.B. Sigman, B.A. Korgel, Electrical transport,Meyer–Neldel rule and oxygen sensitivity
of Bi2S3 nanowires, Nanotechnology 16 (2005) S508.
[4] H. Bao, C.M. Li, X. Cui, Q. Song, H. Yang, J. Guo, Single-crystalline Bi2S3 nanowire network film and
its optical switches, Nanotechnology 19 (2008) 335302.
[5] M.J. Zheng, L.D Zhang, G.H. Li, W.Z. Shen, Fabrication and optical properties of large-scale uniform
zinc oxide nanowire arrays by one-step electrochemical deposition technique, Chem. Phys. Lett. 363
(2002) 123.
[6] T.A. Crowley, K.J. Ziegler, D.M. Lyons, D. Erts, H. Olin, M.A. Morris, J.D. Holmes, Synthesis of metal
and metal oxide nanowire and nanotube arrays within a mesoporous silica template, Chem. Mater. 15
(2003) 3518.
[7] T. Shimizu, T. Xie, J. Nishikawa, S. Shingubara, S. Senz, U. Gösele, Synthesis of vertical high-density
epitaxial Si(100) nanowire arrays on a Si(100) substrate using an anodic aluminum oxide template, Adv.
Mater. 19 (2007) 917.
[8] L. Liu, W. Lee, Z. Huang, R. Scholz, U. Gösele, Fabrication and characterization of a flow-through
nanoporous gold nanowire/AAO composite membrane, Nanotechnology 19 (2008) 335604.
[9] L.P. Barrett, L.G. Joyner, P.P. Halenda, The determination of pore volume and area distributions in
porous substances, J. Am. Chem. Soc. 73 (1951) 373.
[10] P. Birjukovs, N. Petkov, J. Xu, J. Švirksts, J.J. Boland, J.D. Holmes, D. Erts, Electrical characterization
of bismuth sulfide nanowire arrays by conductive atomic force microscopy, J. Phys. Chem. C 112 (2008)
19680.
[11] D. Erts, B. Polyakov, B. Daly, M.A. Morris, S. Ellingboe, J. Boland, J.D. Holmes, High density
germanium nanowire assemblies: contact challenges and electrical characterization, J. Phys. Chem. B 110
(2006) 820.
[12] K. Wang, P. Birjukovs, D. Erts, R. Phelan, M.A. Morris, H. Zhou, J.D. Holmes, Synthesis and
characterisation of ordered arrays of mesoporous carbon nanofibre, J. Mater. Chem. 19 (2009) 1331.
[13] N. Petkov, P. Birjukovs, R. Phelan, M.A. Morris, D. Erts, J.D. Holmes, Growth of ordered
arrangements of one-dimensional germanium nanostructures with controllable crystallinities, Chem.
Mater. 20 (2008) 1902.
[14] M.E. Orazem, B. Tribollet, ECS-The Electrochemical Society 159 (2008) 533 p.
[15] F. Zhang, C. Yuan, X. Lu, L. Zhang, Q. Che, X. Zhang, Facile growth of mesoporous Co3O4 nanowire
arrays on Ni foam for high performance electrochemical capacitors, J. of Power Sources 203 (201) 250.
[16] X. Gan, X. Li, X. Gao, F. Zhuge, W. Yu, ZnO nanowire/TiO2 nanoparticle photoanodes prepared by the
ultrasonic irradiation assisted dip-coating method, Thin Solid Films 518 (2010) 4809.
[17] C.W. Kung, H.W. Chen, C.Y. Lin, K.C. Huang, R. Vittal, K.C. Ho, CoS acicular nanorod arrays for the
counter electrode of an efficient dye-sensitized solar cell, ACS Nano 6 (2012) 7016.
[18] G. Sharma, M.V. Pishko, C.A. Grimes, Fabrication of metallic nanowire arrays by electrodeposition
into nanoporous alumina membranes: effect of barrier layer, J. Mater. Sci. 42 (2007) 4738.
[19] X.Y. Zhang, D. Li, L. Bourgeois, H.T. Wang, P.A. Webley, Direct electrodeposition of porous gold
nanowire arrays for biosensing applications, Chem. Phys. Chem. 10 (2009) 436.
[20] C.H. Liang, K. Terabe, T. Hasegawa, M. Aono, N. Iyi, Anomalous phase transition and ionic
conductivity of AgI nanowire grown using porous alumina template, J. Appl. Phys. 102 (2007) 124308.
[21] J. Xu, N. Petkov, X. Wu, D. Iacopino, A.J. Quinn, G. Redmond, T. Bein, M.A. Morris, J.D. Holmes,
Oriented growth of single-crystalline Bi2S3 nanowire arrays, Chem. Phys. Chem. 8 (2007) 235.
[22] R. Sabino, D. S. Azambuja, R. S. Gonçalves, Electrochemical behavior of aluminum alloy AA2024 in
aqueous solutions in the presence of caffeine, J. Solid State Electrochem.14 (2010) 1255.
[23] IUPAC Recommendations, Impedances of electrochemical systems: terminology, nomenclature and
representation, Pure Appl Chem. 66 (1994) 1831.
[24] A. Lasia, Electrochemical Impedance Spectroscopy and its Applications., Modern Aspects of
Electrochemistry, B. E. Conway, J. Bockris, and R.E. White, Edts., Kluwer Academic/Plenum Publishers,
New York (1999) 32, 143.
6
[25] Release notes for General purose electrochemical system (GPES) for Windows – version 4.9.007 and
Frequency response analysis (FRA) for Windows – version 4.9.007. 2007 Eco Chemie:
www.ecochemie.nl
[26] K.A.Z. Abuthahir, R. Jagannathan, Reverse-loop impedance profile in Bi2S3 quantum dots, Mater.
Chem. Phys. 121 (2010) 184.
[27] B.A. Boukamp, Electrochemical Impedance Spectroscopy, Nano-Electrocatalysis, U. Leiden, 24-28
Nov. (2008) EIS, 61. p.
Figure captions
Fig. 1 Schematic representation of the electrochemical cells with a small volume (1-2 ml) of
electrolyte for electrochemical impedance measurements: 1 − anode (Pt plate); 2 − electrolyte; 3 −
Bi2S3/AAO sample; 4 − Au layer; 5 – Ag/AgCl microelectrode; 6 – contact; 7 – Pt wire.
Fig. 2 Comprehensive surface analysis of the Bi2S3/AAO nanowire arrays: SEM images of (a) low
(77 %) and (b) high (99 %) filling densities of Bi2S3 nanowires in AAO membranes; AFM image of
high density a Bi2S3/AAO sample: (c) topography and (d) current mapping; (e) EDS map of high
density Bi2S3/AAO sample (Al Kα 1.486 eV, Bi Lα 10.839 eV, S Kα 2.309 eV).
Fig. 3 (a) Cyclic voltammetric curves after 30 min immersion in the electrolyte: 1 – AAO
membrane; 2 – low density Bi2S3/AAO nanowire array (77% pore filling); Tafel slope analysis for
(b) high density Bi2S3/AAO array and (c) empty AAO membrane after 30 min of immersion in 0.1
M CH3COONa.
Fig. 4 Nyquist diagram for Bi2S3/AAO nanowire arrays and AAO nanopore sample after 30 min
immersion in 0.1 M CH3COONa at frequency interval 10 kHz – 0.02 Hz and different applied
potentials: a –-0.10 V; b –-0.20 V; 1 – AAO sample; 2 – low density Bi2S3/AAO nanowire sample;
3 – high density Bi2S3/AAO nanowire sample.
Fig. 5 Bode diagram for the empty AAO and Bi2S3/AAO nanowire array samples after 30 min
immersion in aqueous acetate (0.10 mol·L-1), applied potential scan -0.20 V: 1, 4, 7, 8 – AAO
sample; 2, 5 – low density Bi2S3/AAO nanowire sample; 3, 6, 9, 10 – high density Bi2S3/AAO
nanowire samples; 1, 2, 3 – modulus impedance; 4, 5, 6 – phase angle; 7, 10 – real impedance part;
8, 9 – imaginary impedance part.
Fig. 6 Equivalent circuits for electrochemical impedance elements calculation: (a) model for the
empty AAO membrane, R(RQ), (b, c) model for the Bi2S3/AAO nanowire array samples,
Re([R2(R3Q2)]C1).
7
Download